Next Article in Journal
In Silico ADME and Toxicity Prediction of Benzimidazole Derivatives and Its Cobalt Coordination Compounds. Synthesis, Characterization and Crystal Structure
Previous Article in Journal
Triptonoterpene, a Natural Product from Celastrus orbiculatus Thunb, Has Biological Activity against the Metastasis of Gastric Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Future Prospective of Radiopharmaceuticals from Natural Compounds Using Iodine Radioisotopes as Theranostic Agents

by
Wiwit Nurhidayah
1,2,
Luthfi Utami Setyawati
2,3,
Isti Daruwati
3,4,
Amirah Mohd Gazzali
5,
Toto Subroto
1 and
Muchtaridi Muchtaridi
2,3,*
1
Department of Chemistry, Faculty of Mathematics and Natural Sciences, Padjadjaran University, Sumedang 45363, Indonesia
2
Department of Pharmaceutical Analysis and Medicinal Chemistry, Faculty of Pharmacy, Padjadjaran University, Sumedang 45363, Indonesia
3
Research Collaboration Centre for Theranostic Radiopharmaceuticals, National Research and Innovation Agency (BRIN), Sumedang 45363, Indonesia
4
Research Center for Radioisotope, Radiopharmaceutical, and Biodosimetry Technology, Research Organization for Nuclear Energy, National Research and Innovation Agency (BRIN), Serpong 15310, Indonesia
5
School of Pharmaceutical Sciences, Universiti Sains Malaysia, USM, Penang 11800, Malaysia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(22), 8009; https://doi.org/10.3390/molecules27228009
Submission received: 6 September 2022 / Revised: 6 November 2022 / Accepted: 11 November 2022 / Published: 18 November 2022
(This article belongs to the Special Issue Radiolabeled Compounds for Diagnosis and Treatment of Cancer II)

Abstract

:
Natural compounds provide precursors with various pharmacological activities and play an important role in discovering new chemical entities, including radiopharmaceuticals. In the development of new radiopharmaceuticals, iodine radioisotopes are widely used and interact with complex compounds including natural products. However, the development of radiopharmaceuticals from natural compounds with iodine radioisotopes has not been widely explored. This review summarizes the development of radiopharmaceuticals from natural compounds using iodine radioisotopes in the last 10 years, as well as discusses the challenges and strategies to improve future discovery of radiopharmaceuticals from natural resources. Literature research was conducted via PubMed, from which 32 research articles related to the development of natural compounds labeled with iodine radioisotopes were reported. From the literature, the challenges in developing radiopharmaceuticals from natural compounds were the purity and biodistribution. Despite the challenges, the development of radiopharmaceuticals from natural compounds is a golden opportunity for nuclear medicine advancement.

1. Introduction

Radiopharmaceuticals are drugs (pharmaceutical agents) that labeled with radioactive. They could be applied as theranostic agents. Radiopharmaceuticals are required to be target-specific, safe, and effective [1,2]. A radionuclide used for diagnostic purposes usually emits gamma rays, for example, technetium-99m, zirconium-89, indium-111, fluorine-18, xenon-133, iodine-123, and iodine-125. A radionuclide for therapeutic purposes emits alpha or beta rays, such as yttrium-90, iodine-131, samarium-153, lutetium-177, and astatine-211 [3,4]. In addition, the selection of radionuclides also considers half-life, energy, toxicity, and availability in nature. For example, 111In has the ideal SPECT imaging properties but has high cell DNA toxicity. Another example is zirconium-89, whose low pharmacokinetic properties are a perfect radionuclide for antibody labeling. However, the disadvantage is it can provide increased energy and penetrating photons during high abundance production [3].
A pharmaceutical agent can deliver a radiopharmaceutical to a target due to its specific and selective affinity for the target enzyme, protein, or receptor. The consideration in selecting a pharmaceutical agent is the ability to maintain its target specificity and selectivity after radiolabeling [5], and some of the common examples include small molecules such as NaI, peptides, and proteins. However, they still have various stability, specificity, and selectivity problems. These problems prompt the research for new pharmaceutical agents and among the class of compounds with the potential to be developed as natural-based compounds. Natural compounds are known as precursors with various pharmacological effects such as antioxidants, antibacterial, and anticancer activities. The affinity of natural compounds for disease targeting is also favorable in their development as pharmaceutical agents. To achieve this aim, suitable synthesis reactions that will allow stable binding of radionuclides to the natural compounds chosen are needed [6].
One of the radionuclides that are predicted to bind well with natural compounds is iodine isotopes [7]. Iodine has several isotopes, including iodine-123, iodine-124, iodine-125, and iodine-131 (Table 1). The labeling of natural compounds with iodine isotopes in the discovery of radiopharmaceuticals for diagnostics and therapeutics of various diseases is very promising. Natural compounds that act as pharmaceutical agents will deliver iodine isotopes to the target. The radiation emitted by the iodine isotopes will then function as either a diagnostic or a therapeutic agent (Figure 1).
Although very promising, the development of radiopharmaceuticals from natural compounds using iodine isotopes is limited. More information needs to be gathered in order to understand the limitations and challenges of this approach, and to drive effective strategies for the research and development process. This review collects data of radiopharmaceuticals from natural compounds using iodine radioisotopes that were reported in the last ten years. The stages and the challenges were reviewed, and potential strategies were discussed to escalate the development of the radiopharmaceuticals. The information gained in this review will help researchers to advance the research and development of radiopharmaceuticals and nuclear medicine practice in the future.

2. Differences between Radiopharmaceuticals from Natural Compounds with Other Radiopharmaceuticals

Radiopharmaceuticals from natural compounds are radiopharmaceuticals that use natural compounds as the ligand. The ligand will then selectively interact with target tissues; thus, it has the ability to selectively deliver radionuclides. This interaction can occur pharmacologically, immunologically, or metabolically, and they may be reversible. After the interaction and the binding of the ligand with its target, the bonded radiopharmaceutical can be internalized and stored in the target cells. It is hence very crucial for the ligand to effectively, at a low concentration, prevent any pharmacological activity or side effects on the target [5].
In comparison with radiopharmaceuticals from natural compounds, conventional radiopharmaceuticals usually utilize small molecules, peptides, and proteins as ligands. Small molecules such as amino acids, fatty acids, nucleotides, and small inorganic molecules enable the targeting of intracellular regions because small molecules can penetrate semipermeable membranes easily. The examples of radiopharmaceuticals that use small molecules are [123I]NaI, [123I]ioflupane, and [123]] iobenguane for neuroblastoma tumors [5]. [123I]NaI is a substrate for sodium iodide symporter for thyroid imaging. The presence of a parathyroid adenoma is characterized by areas of cellular tissue which do not exhibit trapping of [123I]NaI [16]. [123I] ioflupane provides sensitive results in the diagnosis of Parkinson’s disease even in its early stages, based on the pattern of [123I]ioflupane uptake on SPECT images, which can be interpreted as normal activity if it shows no dopaminergic deficit [17]. [123I] iobenguana was applied to neuroblastoma tumors by targeting the norepinephrine transporter (NET) [18].
Radiopharmaceuticals that use peptides or proteins usually target specific receptors of tumor or cancer cells. Peptide cells can diffuse rapidly into target tissues and show longer accumulation in tumor cells. However, the disadvantage of using peptides or proteins as ligands is the potential for radio nephrotoxicity due to the high accumulation of peptides in the kidney. One example of a protein as a ligand is the Designed ankyrin repeat proteins (DARPins) labeled with iodine-124, iodine-125 and iodine-131 which is aimed to evaluate human epidermal growth factor receptor 2 (HER2) expression levels in breast and gastroesophageal cancer [19].
The difference between radiopharmaceuticals from natural compounds and other radiopharmaceuticals is also found in the stages of development of these drugs as shown in Figure 2. In general, the stages of development of other radiopharmaceuticals consist of identifying the molecular targets and synthesizing pharmaceutical compounds to be used as ligands (small molecules or peptides). Suitable radiopharmaceutical synthesis reaction will then be selected, and the evaluation of the synthesized radiopharmaceutical will be carried out. Meanwhile, radiopharmaceuticals from natural compounds tend to have a longer development stage. The initial stage of the development of radiopharmaceuticals from natural compounds is the discovery of natural compounds themselves. Research usually starts from exploring the sources of natural products in nature. After that, the lead compound will be identified, and the natural compounds will be isolated and identified by their structure elucidation. Subsequently, the molecular targets and pharmacological activities will be identified [6,20]. The next step is the selection of an appropriate radiopharmaceutical synthesis reaction based on the structure and the targets of the natural compounds. Some natural compounds also require structure modifications to get the best radiopharmaceutical synthesis results. Then, just like other radiopharmaceuticals, the radiopharmaceuticals from natural compound will also be characterized and evaluated based on the following criteria; the stability, physicochemical characteristics, cellular uptake, preclinical studies, dosimetry prediction, and clinical studies [6].

3. Available Literature on from Natural Compounds with Iodine Radioisotopes in Last 10-Year Period

This review summarizes the reported studies on radiopharmaceuticals from natural compounds with iodine radioisotopes in the last ten years, between 2013 and 2022. The data obtained from various research are presented in Figure 3.
From 2013–2022, 32 analyses on radiopharmaceuticals from natural compounds with iodine radioisotopes were conducted. The natural compounds are mostly isolated from plants, and their pharmacological effects are identified. These natural compounds are labeled with iodine radioisotopes to develop several therapeutic or diagnostic agents for various diseases, including nineteen for tumor or cancer, one for urinary tract dysfunction, one for Alzheimer’s disease, seven for necrotic myocardium, one for neuroblastoma, one for ischemic stroke, one for determination of natural compound toxicity, and another one was just labelled as radiotracer unspecified. As described in Figure 2, the discovery and development of radiopharmaceuticals have a long process. After conducting radiolabeling reaction or radiopharmaceutical synthesis, several evaluations need to be passed, including stability, physicochemical, cellular uptake, preclinical, dosimetry, and clinical studies. From the recent research based on the collected data from the 32 studies, 1 was in the synthesis stage, 1 was in the physicochemical study stage, 7 were in the cellular uptake study stage, 21 were in the preclinical study stage, 2 were in the dosimetry prediction stage and none of them reached the clinical study stage. Detailed information is listed in Table 2.

4. Synthesis of Radiopharmaceuticals from Natural Compounds with Iodine Radioisotopes

In the development of radiopharmaceuticals, researchers commonly favor two kinds of synthesis reactions: (1) synthesis with nonradioactive iodine (iodine-127), and (2) radiosynthesis with iodine radioisotope. The purpose of nonradioactive synthesis is to predict the structure of the objective compound by elucidating the structure using MS (Mass Spectroscopy) and NMR (Nuclear Magnetic Resonance). Radiosynthesis was to be tested for radiochemical purity. They are usually carried out through two reaction mechanisms, namely (1) electrophilic substitution and (2) nucleophilic substitution [7,93].

4.1. Electrophilic Substitutions

Electrophilic substitution reactions occur when iodine substitutes hydrogen on electron-rich aromatic rings such as phenol and group-substituted benzene rings. In general, iodine is available in the form of NaI solution therefore it needs to be converted into an electropositive form before reacting with pharmaceutical compounds using oxidizing agents, iodo–deprotonation, and iodo–demetallation. The oxidizing agent converts iodine to its electropositive form by oxidizing it thus its oxidation number increases. There are two types of oxidizing agent: (1) oxidizing agents containing halogens, and (2) oxidizing agents without halogens. Oxidizing agents with halogens include chloramine-T, iodine, and N-chlorosuccinimide, meanwhileoxidizing agents without halogens include tert-butyl hydroperoxide, peracetic acid and hydrogen peroxide [94]. Iododeprotonation usually occurs in aromatic compounds that have an electron-rich ring activated by OH, NH2, or OMe [95]. Iododemetallization is reaction using organometallic precursors such as trialkylstannyl, trialkylsilyl, or boronic acid derivatives [93].

4.2. Nucleophilic Substitution

Nucleophilic substitution reactions consist of several methods, including halogen exchange, isotope exchange, radioiodo-dediazonisation, and copper-assisted halogen exchange. The halogen exchange method occurs when radioactive iodine substitutes a halogen (bromine or chlorine) in the pharmaceutical compound [93]. This reaction requires extreme conditions. Zmuda et al. (2015) synthesized a tracer for Poly (ADP-ribose) Polymerase-1 (PARP-1) with solid state halogen exchange radioiodination method using bromination. The reaction took place under extreme conditions whereby the reaction temperature was 210 °C with an incubation time of 0.5 h. The authors reported the radiochemical yield obtained was 36.5 ± 7.2% [93,96].
The isotope exchange reaction was carried out by substituting the iodine present in the ligand with iodine radioactive. This reaction usually occurs under reflux with solvents. The solvents used are acetone, dichloromethane, acetonitrile, water, ethanol, or methyl ethyl ketone. Sadeghzadeh et al. synthesized 4-benzyl-1-(3-[125I]iodobenzylsulfonyl)piperidine and 4-(3-[125I]iodobenzyl)-1-(benzylsulfonyl)piperazine using this reaction. It used a wet method using different organic solvents, such as propylene glycol at elevated temperatures (100–200 °C) where the results showed that the purity obtained was 70% [93,97]. Radioiodo–dediazonisation is a radioiodination method of compounds with a diazonium group. The reaction was conducted by substituting diazonium with radioiodine. It is usually carried out at low temperatures with the help of sodium nitrate. The reaction proceeds by the SN1 mechanism [93]. Copper-assisted exchange is a nucleophilic substitution reaction that uses copper as a catalyst. The reaction can occur via isotopic or halogen exchange. M. Hagimori et al. synthesized matrix metalloproteinase-12 (MMP-12) using a copper-assisted exchange. The reaction was conducted at 140 °C for 60 min, with high purity product [93,98].

4.3. Synthesis of Radiopharmaceuticals from Natural Compound with Iodine Radioisotopes in the Last 10 Years

The synthesis of radiopharmaceuticals from natural compounds is a challenging process. Natural compounds are expected to be labeled as stable and produce high radiochemical purity. The natural compounds to be labeled are mostly isolated from plants. Prior to labeling, they are usually characterized by LC/MS, HPLC, or NMR. Based on the literature study, several research articles reported the characterization method of natural compounds, but several articles have not reported it. Out of the 32 radiopharmaceuticals of natural compounds using iodine radioisotope, 17 reported the natural compound characterization method, while 15 have not reported it. From the data collected in the last 10 years, all reported compounds were synthesized through electrophilic substitution reactions. This is because natural compounds would usually have an electron-rich aromatic ring. Of the 32 labeled compounds, 31 were synthesized in the presence of oxidizing agents, and 1 through iodo–destannylation. As oxidizing agents, 20 compounds used iodogen, 9 with chloramine-T, and 2 compounds with peracetic acids. Detailed information is listed in Table 3.
The synthesis method and type of oxidizing agent used in the synthesis, and the purity of the 32 radiopharmaceutical candidates are described in Figure 4. The results of radiosynthesis showed that 24 compounds (77%) had radiochemical purity above 95% while 8 (23%) had a purity lower than 95%.

5. Evaluations of Radiopharmaceuticals from Natural Compounds with Iodine Radioisotopes

As illustrated in Figure 2, the evaluation of radiopharmaceuticals includes stability tests, physicochemical analysis, cellular uptake study, preclinical study, dosimetry prediction, and clinical study. All radiopharmaceuticals, including those derived from natural compounds, will need to go through these evaluations before approval can be granted for human use. Table 4 presents the evaluation of the 32 radioiodinated natural compound as collected and analyzed for this review. From the data, 22 of them reported stability tests, 8 of them reported physicochemical analysis, 10 of them reported cellular uptake study, 23 of them reported preclinical study, 2 of them reported dosimetry prediction and non-reported clinical study.
The stability of radiopharmaceuticals is affected by several factors such as pH, light, and temperature. It needs to be stored under various storage conditions [98]. From 22 radioiodinated natural compounds that reported stability tests, the stability was classified into two groups: stability ≥ 24 h and <24 h, with a total of 14 and 8 compounds, respectively.
The physicochemical analysis consists of lipophilicity and protein binding characterizations. Lipophilicity, quantified as Log D or Log P, is a crucial parameter in estimating radiopharmaceuticals absorption, distribution, metabolism, and excretion (ADME) [99,100]. Besides lipophilicity, protein binding affects the biodistribution and clearance of radiopharmaceuticals. It has a positive correlation with lipophilicity [101].
Cellular uptake study aims to determine the specificity of a radiopharmaceutical towards its target by using cells or tissues that expresses the target [4,102]. The tested radiopharmaceutical will be incubated with cultured cells and cell uptake will be calculated as the percentage of radioactivity in cells compared to total radioactivity [103]. The selection of cell lines used depends on the specific target of the radiopharmaceutical on the receptor or on certain physiological conditions. Some of the cell lines that are often used include: Hutu80 (human gastrointestinal tumor cell lines), Caco-2 (human colon adenocarcinoma cells), MCF7 (human breast adenocarcinoma cells), PC3 (human prostate carcinoma cells), Keratinocyte (Human normal epidermal keratinocyte cells), BJ (Human normal foreskin fibroblast cells), TT, FTC-133, and DRO (human thyroid cell lines), SK-N-AS and SH-SY5Y (human neuroblastoma cell lines), MDA-MB-231 (the triple-negative breast cancer cell lines), and SKOV3 (human ovarian cancer cell lines). After cellular uptake, preclinical study will be conducted using experimental animals. In general, it consists of biodistribution, pharmacokinetics, and toxicity studies. A biodistribution study will allow the determination of radiopharmaceutical uptake in the animal organs, which will be calculated as %ID/g [104]. Based on the reported biodistribution study data, several compounds ([131I]hydroxytyrosol, [123I]hesperetin, [125I]rutin, [131I]khellin, and [125I]zearalenone) have a high accumulation pattern in certain organs, especially the thyroid, intestine and stomach. A pharmacokinetic study is needed to determine the pharmacokinetic parameters such as elimination rate constant (Ke), the volume of distribution (Vd), area under the curve [105], and clearance and time to maximum concentration (Tmax) [106]. Radiopharmaceuticals are expected to have rapid blood clearance and short t1/2 elimination so that they can be excreted rapidly from the blood. Based on data collected, [131I]sennidin A, [131I]protohypericin, [131I]sennoside B, [131I]rhein, [131I]vitexin, [131I]napthazarin, and [131I]shikonin reported pharmacokinetic study with t1/2 elimination value of 11.75, 14.9, 8.6, 8.2, 5.3, 4.73, and 0.675 h, respectively. Toxicity study aims to evaluate the safety of radiopharmaceuticals. Koziorowski et al. (2016) stated that in radiopharmaceuticals, acute toxicity tests were carried out to predict the effect of overdose whereas subacute, chronic, teratogenic, mutagenic and carcinogenic toxicity were not required for radiopharmaceuticals [107]. The schematic diagram of the preclinical study is depicted in Figure 5.
Dosimetry prediction is a procedure in the determination of the absorbed dose as the amount of energy absorbed per unit mass in all irradiated tissues or organs of interest. The aim was to determine the reference levels of irradiation for every new radiopharmaceutical or estimate the absorbed dose for routinely used radiopharmaceuticals [108]. The final stage of radiopharmaceutical development is clinical study. This stage is carried out based on the regulations set in each region or country because, in general, each region or country would have different regulations regarding the rules of radiopharmaceutical-based clinical trials [5].

6. Challenge and Strategies

The development stages of radiopharmaceuticals from natural compounds with iodine radioisotope are rather long and challenging. The challenges reported in previous studies were mostly related to radiochemical purity and biodistribution, as shown in Table 5. However, some radiolabeled compounds such as [131I]genistein have not been tested in vitro and in vivo, so further development is needed.

6.1. Problem Related to Radiochemical Purity and the Strategies

Radiochemical purity is a crucial quality control factor in radiopharmaceutical development. The radiochemical impurities of iodine are free I and I2 which affect the safety and accuracy of radiopharmaceuticals. The first strategy to obtain the maximum radiochemical purity is to select a suitable radioiodination method based on the steric characteristics of the natural compound as a substrate to be labeled. In addition, critical point optimization in the radioiodination method is important because it can minimize the formation of impurities. The first strategy is to optimize the critical point of radioiodination. [7,94]. Substrate characteristics and critical reaction points should be considered in selecting the radioiodination reaction method, as summarized in Table 6.
Selection of the suitable radioiodination method and optimization of the critical point in radioiodination could lead to high radiochemical purity. However, if the radiochemical purity is still lower than the required purity (>95%), another strategy that can be conducted is purification to separate impurities from the radiopharmaceuticals. The selection of the purification method depends on the molecular weight, lipophilicity, and molecular charge of the radiopharmaceuticals. Purification methods that can be applied include HPLC (High-Performance Liquid Chromatography), SPE (Solid Phase Extraction), SEC (Size-Exclusion Chromatography), and IEC (Ion-Exchange Chromatography.
One of the most widely used purification methods is the HPLC. The separation of compounds occurs due to differences in solute interactions and the column that lead to different elution rates for each component. As a result, it will provide high purity resolution. The parameters to consider in HPLC purification are polarity, flow rate, pH, the lipophilicity of the mobile phase, sample matrix, type of stationary phase, and temperature. SPE is widely chosen because it is simple, fast, and able to separate dissolved or suspended compounds from other compounds in the mixture based on their physical and chemical properties. Kim et al. (2019) conducted purification of [131I]metaiodobenzylguanidine using solid phase extractionand obtained a higher amount of product and lower exposure of operator to radiation [122]. SPE is commonly used in the separation of macromolecules that consist of substances with different molar masses. The SEC chromatography column uses porous polymeric beads. The pore size determines the dimensions of the compounds to be separated. Molecules with smaller size than the pores can enter the pores and retain, while the molecule with larger size than the pores will pass through the spaces between the packing material. In this way, the molecules with the highest molecular weight will be obtained in the first fraction. Lemps et al. performed purification of [123I]bevacizumab using the SEC method and obtained a radiochemical purity of 99.5% after purification [123]. IEC is a separation method for ions and polar molecules [124]. IEC consists of anion and cation exchange. Cation-exchange chromatography uses a negatively charged stationary phase that can separate cations from other ions. By contrast, anion-exchange chromatography uses a positively charged stationary phase that can separate anions from the other ions. Before conducting an IEC, the stationary phase should be achieved through electroneutrality. Visser et al. performed a purification of [131I]c-MOv18 which was a radiopharmaceutical candidate for therapy for ovarian cancer. The impurities were removed by purification using Dowex AG1-X8 (BioRad, Utrecht, The Netherlands) anion-exchange resin in PBS [125,126].

6.2. Problem Related to Biodistribution and the Strategies

Ideally, radiopharmaceuticals are required to have high specificity, rapid accumulation in target organs, and a high target-to-nontarget ratio [127]. Based on previous studies, some labeled compounds showed high biodistribution in other organs compared to the targeted organs. Altered biodistribution in vivo affected the accuracy of imaging and radiopharmaceutical therapy. This problem occurred due to the presence of other compounds, such as impurities (free I) or residues that could be uptaken in organs and detected as radiopharmaceuticals. Spetz et al. conducted a study to determine the biodistribution of free 125I and 131I in rats and reported the highest biodistribution in the thyroid gland and stomach. In primary conditions, iodine was localized in the thyroid. In addition, the iodine uptake in the stomach occurred due to the expression of an iodine transport medium named Na+/I− sodium iodide symporter in the stomach [128].
The formation of impurity free iodide indicates low in vivo stability of C-I in the radiopharmaceuticals due to the deiodination reaction. The accumulation of free iodide as impurity in the thyroid, stomach, and intestine reduces the target-to-background ratio of the diagnostic agent so that the diagnostic results are biased. In therapeutic applications, it increases the accumulation of radioactivity in the non-target organs, which could lead to adverse reactions in these healthy organs. This in vivo deiodination reaction is caused by several enzymes including deiodinase enzymes, cytochromes P450 (CYP450) enzymes, and nonspecific nucleophilic enzymes. Deiodinase enzymes promote deiodination reactions in iodinated aromatic rings such as ortho–iodo–phenols. CYP450 enzymes promote deiodination via xenobiotics oxidation reactions, whereas nonspecific nucleophilic enzymes promote deiodination at electrophilic carbon atoms [129].
The first strategy to decrease free iodine accumulation in non-target organs is to design radiopharmaceuticals resistant to deiodination reactions with modification structural. In general, compounds with an arene group are stable to deiodination. In the iodination of the arene group, metaiodoarene is more resistant to deionization than orthoidoarene and paraiodoarene. In addition, iodination at sp2 carbon atoms is usually more stable to deiodination reactions than iodination at sp and sp3 carbon atoms. Radioiodination of the vinyl group is also stable against in vivo deiodination reactions. However, the phenol and aniline groups have poor in vivo stability. Their stability can be improved by adding electron-donating substituents such as OCH3 to the aromatic ring. On the other hand, the addition of electron-withdrawing groups can decrease in vivo stability [129]. The resistance of some radioiodinated groups against in vivo deiodination is listed in Table 7.
Compton et al. (1993) conducted the radioiodination of ∆9-Tetrahydrocannabiol (∆9-THC) to produce 2-iodo-∆8-THC and 5′-iodo-∆8-THC. ∆9-THC is a natural compound isolated from Cannabis Sativa. The target of radioiodinated ∆9-THC is the imaging cannabinoid system. The structure of ∆9-THC, 2-iodo-∆8-THC, and 5′-iodo-∆8-THC are shown in Figure 6a. The in vivo study of 5′-iodo-∆8-THC showed a poor in vivo profile due to the deiodination reaction (shown in Figure 6b). The position of iodine on a terminal sp3 carbon atom of linear pentyl moiety allows 5′-iodo-∆8-THC to be susceptible to in vivo deiodination caused by CYP450. Cavina et al. (2016) provide solutions for structural modifications that are expected to increase the stability against deiodination, including the position of iodine on the iodo-ethoxy group, iodine on the cubane position, iodine on carbon sp2 at the vinyl terminal, or iodine on C sp2 in the terminal allyl moiety [129]. The structural modification is shown in Figure 6c. Based on this case, structural modification of natural compounds provides a promising strategy to increase the stability of radioiodinated natural compounds against in vivo deiodination.
The second strategy to increase the in vivo stability of natural compound-based radiopharmaceutical candidates is to label them with a linker. This linker will form a stable chemical bond between the iodine radioisotope and the natural compounds [5], and it must have a good in vivo stability. Kim et al. (2016) conducted radioiodination of cetuximab with the linker (N-(4-isothiocyanatobenzyl)-2-(3-(tributylstannyl)phenyl) acetamide (IBPA). It was reported that [125I]IBPA-cetuximab had a more stable binding and higher internalization in mice bearing LS174T tumor xenografts compared to [125I]cetuximab [130].
Another strategy that can be applied is to increase the target specificity by using nanoparticles. Nanoparticles play a role in increasing the penetration of compounds across biological membranes so that they can effectively deliver therapeutic agents and reduce the side effects of conventional delivery techniques. Nanoparticles can increase the delivery specificity of radiopharmaceuticals towards its target by conjugating the targeting molecule (ligand) on the nanoparticles’ surface [131]. This technique was carried out by Ince et al. (2016) who produced the [131I]FATQCSNPs (folic acid-chitosan nanoparticles loaded with thymoquinone) described earlier, with ovarian cancer cells as the target. [131I]FATQCSNPs incorporated thymoquinone isolated from Nigella sativa in a folic acid-chitosan nanoparticles [74]. Folic acid is a small molecule that is useful as a ligand that helps in the internalization of pharmaceuticals into cancer cells [75]. Folic acid was used due to ovarian cancer cells that were marked with overexpression of folic acid. The encapsulation of thymoquinone within folic acid-chitosan nanoparticles has improved the delivery of thymoquinone, especially with the presence of folic acid that helps to increase the specificity of delivery and increase cellular uptake. Both [131I]thymoquinone and [131I]FATQSCNPs were developed for diagnostic and therapy of cancer. The radioiodinated complex showed a higher uptake in SKOV3 cells as compared to [131I]thymoquinone [76]. A schematic diagram of [131I]FATQSCNPs is shown in Figure 7.

7. Methods

This review was conducted based on the results of the collection and analysis of articles obtained from the PubMed database with the following keywords: “radioiodination of natural compound”; “radiopharmaceutical natural compound”; “radioiodination of flavonoid”; “radioiodination of alkaloid”; “radioiodination reaction mechanism”; “design AND challenge new radiopharmaceutical”; “radiopharmaceutical AND natural product AND iodine”.
The inclusion criteria of the main article were articles that discuss radioiodination of natural compounds using the English language and published within the range years of 2013-2022. The inclusion criteria of the supporting articles were articles that discuss radiopharmaceuticals in general, iodine isotopes, and the pharmacological effects of natural compounds. Exclusion criteria were articles that were published more than 10 years ago for main articles, 20 years ago for supporting articles, and non-relevant articles to the topic discussed in this review.
Based on the search conducted using the aforementioned keywords, 942 journals were obtained and 512 articles were discarded as they were published more than 10 years ago (for main articles) and 20 years ago (for supporting articles), while 299 articles were non-relevant articles to the topic discussed in this review. This step has reduced the number of articles to 131 consisting of 102 supporting articles and 29 articles discussing the radioiodination of natural compounds with iodine radioisotope. The literature search flow is shown in Figure 8.

8. Future, Prospect, and Conclusions

The development of radiopharmaceuticals from natural compounds with iodine radioisotope is a long process with several challenges. Thirty-two radioiodinated natural compounds were collected from a literature study of the last 10 years. To determine the challenges that radiopharmaceutical researchers found in natural compounds, we reviewed 32 compounds from their synthesis to their evaluation results. These challenges are clasified into two groups: (1) challenges related to chemical purity, and (2) challenges related to biodistribution. We discussed strategies that could be applied to resolve these challenges.
Based on the data, 8 of the 32 radioiodinated natural compounds collected had radiochemical purity problems. The first strategy offered is to optimize the critical point in the synthesis reaction to obtain the optimum synthesis conditions. The second strategy is the purification of synthetic products in several ways, including High-Performance Liquid Chromatography (HPLC), SPE (Solid Phase Extraction), (SPE), SEC (Size-Exclusion Chromatography), and IEC (Ion-Exchange Chromatography). The purification method can separate the product from impurities.
Based on the evaluation results, five radioiodinated natural compounds have problems with their biodistribution. Unspecified accumulation is characterized by high accumulation in the stomach, intestines, and thyroid. This unspecific accumulation shows poor in vivo stability due to the deiodination reaction. Several strategies to solve this problem include designing radiopharmaceuticals resistant to in vivo deiodination by structural modification, radioiodination with linkers, and application of nanoparticles.
Despite the challenges, the development of radiopharmaceuticals from natural compounds using iodine radioisotope offers a bright future in the development of radiopharmaceuticals. This review provides information that researchers undertaking further research can consider. The strategies offered in this review are expected to encourage improvement in research related to natural product-based radiopharmaceuticals with iodine radioisotopes as theranostic agents for various diseases.

Author Contributions

W.N. and M.M. contributed to design the concept and the content of manuscript, literature search, and manuscript preparation. I.D., L.U.S. and A.M.G. contributed to manuscript review. T.S. and M.M. contributed to design the concept and the content of manuscript, manuscript review, and M.M. responsible as guarantor. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by PMDSU Scholarship, grant number 2064/UN6.3.1/PT.00/2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank the Ministry of Education and Culture of Republic of Indonesia for financially support the review through PMDSU scholarship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sgouros, G.; Bodei, L.; McDevitt, M.R.; Nedrow, J.R. Radiopharmaceutical therapy in cancer: Clinical advances and challenges. Nat. Rev. Drug Discov. 2020, 19, 589–608. [Google Scholar] [CrossRef] [PubMed]
  2. Lau, J.; Rousseau, E.; Kwon, D.; Lin, K.-S.; Bénard, F.; Chen, X. Insight into the Development of PET Radiopharmaceuticals for Oncology. Cancers 2020, 12, 1312. [Google Scholar] [CrossRef] [PubMed]
  3. Holik, H.A.; Ibrahim, F.M.; Elaine, A.A.; Putra, B.D.; Achmad, A.; Kartamihardja, A.H.S. The Chemical Scaffold of Theranostic Radiopharmaceuticals: Radionuclide, Bifunctional Chelator, and Pharmacokinetics Modifying Linker. Molecules 2022, 27, 3062. [Google Scholar] [CrossRef] [PubMed]
  4. Payolla, F.; Massabni, A.; Orvig, C. Radiopharmaceuticals for diagnosis in nuclear medicine: A short review. Eclét. Quím. J. 2019, 44, 11–19. [Google Scholar] [CrossRef]
  5. Vermeulen, K.; Vandamme, M.; Bormans, G.; Cleeren, F. Design and Challenges of Radiopharmaceuticals. Semin. Nucl. Med. 2019, 49, 339–356. [Google Scholar] [CrossRef]
  6. Wongso, H. Natural product-based Radiopharmaceuticals:Focus on curcumin and its analogs, flavonoids, and marine peptides. J. Pharm. Anal. 2021, 12, 380–393. [Google Scholar] [CrossRef]
  7. Vaidyanathan, G.; Zalutsky, M.R. The Radiopharmaceutical Chemistry of the Radioisotopes of Iodine. In Radiopharmaceutical Chemistry; Lewis, J.S., Windhorst, A.D., Zeglis, B.M., Eds.; Springer International Publishing: Berlin/Heidelberg, Germany, 2019; pp. 391–408. [Google Scholar] [CrossRef]
  8. Morphis, M.; van Staden, J.A.; du Raan, H.; Ljungberg, M. Validation of a SIMIND Monte Carlo modelled gamma camera for Iodine-123 and Iodine-131 imaging. Heliyon 2021, 7, e07196. [Google Scholar] [CrossRef]
  9. Yordanova, A.; Eppard, E.; Kürpig, S.; Bundschuh, R.A.; Schönberger, S.; Gonzalez-Carmona, M.; Feldmann, G.; Ahmadzadehfar, H.; Essler, M. Theranostics in nuclear medicine practice. OncoTargets Ther. 2017, 10, 4821–4828. [Google Scholar] [CrossRef] [Green Version]
  10. Silberstein, E.B. Radioiodine: The classic theranostic agent. Semin. Nucl. Med. 2012, 42, 164–170. [Google Scholar] [CrossRef]
  11. Treglia, G.; Muoio, B.; Giovanella, L.; Salvatori, M. The role of positron emission tomography and positron emission tomography/computed tomography in thyroid tumours: An overview. Eur. Arch. Oto-Rhino-Laryngol. 2013, 270, 1783–1787. [Google Scholar] [CrossRef]
  12. Braghirolli, A.M.; Waissmann, W.; da Silva, J.B.; dos Santos, G.R. Production of iodine-124 and its applications in nuclear medicine. Appl. Radiat. Isot. 2014, 90, 138–148. [Google Scholar] [CrossRef] [PubMed]
  13. Cascini, G.L.; Asabella, A.N.; Notaristefano, A.; Restuccia, A.; Ferrari, C.; Rubini, D.; Altini, C.; Rubini, G. 124 Iodine: A longer-life positron emitter isotope-new opportunities in molecular imaging. BioMed Res. Int. 2014, 2014, 672094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Schwarz, S.B.; Thon, N.; Nikolajek, K.; Niyazi, M.; Tonn, J.C.; Belka, C.; Kreth, F.W. Iodine-125 brachytherapy for brain tumours—A review. Radiat. Oncol. 2012, 7, 30. [Google Scholar] [CrossRef] [Green Version]
  15. Takiar, V.; Voong, K.R.; Gombos, D.S.; Mourtada, F.; Rechner, L.A.; Lawyer, A.A.; Morrison, W.H.; Garden, A.S.; Beadle, B.M. A choice of radionuclide: Comparative outcomes and toxicity of ruthenium-106 and iodine-125 in the definitive treatment of uveal melanoma. Pr. Radiat. Oncol. 2015, 5, e169–e176. [Google Scholar] [CrossRef] [PubMed]
  16. Pelletier-Galarneau, M.; Sogbein, O.O.; Dinh, L.; Zuckier, L.S. Superiority of Digital Subtraction for Analysis of Simultaneously-Acquired Dual-Radiopharmaceutical Parathyroid Scintigraphy. Open J. Med. Imaging 2015, 5, 42–48. [Google Scholar] [CrossRef] [Green Version]
  17. Prashanth, R.; Roy, S.D.; Mandal, P.K.; Ghosh, S. High-Accuracy Classification of Parkinson’s Disease Through Shape Analysis and Surface Fitting in 123I-Ioflupane SPECT Imaging. IEEE J. Biomed. Health Inform. 2017, 21, 794–802. [Google Scholar] [CrossRef] [Green Version]
  18. Pandit-Taskar, N.; Modak, S. Norepinephrine Transporter as a Target for Imaging and Therapy. J. Nucl. Med. 2017, 58, 39S. [Google Scholar] [CrossRef] [Green Version]
  19. Vorobyeva, A.; Schulga, A.; Konovalova, E.; Güler, R.; Mitran, B.; Garousi, J.; Rinne, S.; Löfblom, J.; Orlova, A.; Deyev, S.; et al. Comparison of tumor-targeting properties of directly and indirectly radioiodinated designed ankyrin repeat protein (DARPin) G3 variants for molecular imaging of HER2. Int. J. Oncol. 2019, 54, 1209–1220. [Google Scholar] [CrossRef] [Green Version]
  20. Hanson, J. A Hundred Years in the Elucidation of the Structures of Natural Products. Sci. Prog. 2017, 100, 63–79. [Google Scholar] [CrossRef]
  21. Chimento, A.; Casaburi, I.; Rosano, C.; Avena, P.; De Luca, A.; Campana, C.; Martire, E.; Santolla, M.F.; Maggiolini, M.; Pezzi, V.; et al. Oleuropein and hydroxytyrosol activate GPER/ GPR30-dependent pathways leading to apoptosis of ER-negative SKBR3 breast cancer cells. Mol. Nutr. Food Res. 2014, 58, 478–489. [Google Scholar] [CrossRef]
  22. Sirianni, R.; Chimento, A.; De Luca, A.; Casaburi, I.; Rizza, P.; Onofrio, A.; Iacopetta, D.; Puoci, F.; Andò, S.; Maggiolini, M.; et al. Oleuropein and hydroxytyrosol inhibit MCF-7 breast cancer cell proliferation interfering with ERK1/2 activation. Mol. Nutr. Amp. Food Res. 2010, 54, 833–840. [Google Scholar] [CrossRef] [PubMed]
  23. Luo, C.; Li, Y.; Wang, H.; Cui, Y.; Feng, Z.; Li, H.; Li, Y.; Wang, Y.; Wurtz, K.; Weber, P.; et al. Hydroxytyrosol promotes superoxide production and defects in autophagy leading to anti-proliferation and apoptosis on human prostate cancer cells. Curr. Cancer Drug Targets 2013, 13, 625–639. [Google Scholar] [CrossRef]
  24. Sun, L.; Luo, C.; Liu, J. Hydroxytyrosol induces apoptosis in human colon cancer cells through ROS generation. Food Funct. 2014, 5, 1909–1914. [Google Scholar] [CrossRef] [PubMed]
  25. Toteda, G.; Lupinacci, S.; Vizza, D.; Bonofiglio, R.; Perri, E.; Bonofiglio, M.; Lofaro, D.; La Russa, A.; Leone, F.; Gigliotti, P.; et al. High doses of hydroxytyrosol induce apoptosis in papillary and follicular thyroid cancer cells. J. Endocrinol. Investig. 2017, 40, 153–162. [Google Scholar] [CrossRef] [PubMed]
  26. Ozkan, M.; Muftuler, F.; Yurt, A.; Medine, I.; Unak, P. Isolation of Hydroxytyrosol from olive leaves extract, radioiodination and investigation of bioaffinity using in vivo/in vitro methods. Radiochim. Acta 2013, 101, 585–593. [Google Scholar] [CrossRef]
  27. Khalil, N.; Bishr, M.; Desouky, S.; Salama, O. Ammi visnaga L., a Potential Medicinal Plant: A Review. Molecules 2020, 25, 301. [Google Scholar] [CrossRef] [Green Version]
  28. Khater, S.I.; Kandil, S.A.; Hussien, H. Preparation of radioiodinated khellin for the urinary tract imaging. J. Radioanal. Nucl. Chem. 2013, 295, 1939–1944. [Google Scholar] [CrossRef]
  29. Mullaicharam, A.R.; Nirmala, H. St John’s wort (Hypericum perforatum L.): A Review of its Chemistry, Pharmacology and Clinical properties. Int. J. Res. Phytochem. Pharmacol. Sci. 2018, 1, 5–11. [Google Scholar] [CrossRef]
  30. Cona, M.M.; Koole, M.; Feng, Y.; Liu, Y.; Verbruggen, A.; Oyen, R.; Ni, Y. Biodistribution and radiation dosimetry of radioiodinated hypericin as a cancer therapeutic. Int. J. Oncol. 2014, 44, 819–829. [Google Scholar] [CrossRef]
  31. Cona, M.M.; Alpizar, Y.A.; Li, J.; Bauwens, M.; Feng, Y.; Sun, Z.; Zhang, J.; Chen, F.; Talavera, K.; de Witte, P.; et al. Radioiodinated Hypericin: Its Biodistribution, Necrosis Avidity and Therapeutic Efficacy are Influenced by Formulation. Pharm. Res. 2014, 31, 278–290. [Google Scholar] [CrossRef]
  32. Pour, A.P.; Farahbakhsh, H. Lawsonia inermis L. leaves aqueous extract as a natural antioxidant and antibacterial product. Nat. Prod. Res. 2020, 34, 3399–3403. [Google Scholar] [CrossRef] [PubMed]
  33. Tekin, V.; Muftuler, F.Z.B.; Yurt Kilcar, A.; Unak, P. Radioiodination and biodistribution of isolated lawsone compound from Lawsonia inermis (henna) leaves extract. J. Radioanal. Nucl. Chem. 2014, 302, 225–232. [Google Scholar] [CrossRef]
  34. Tekin, V.; Biber Muftuler, F.Z.; Guldu, O.K.; Kilcar, A.Y.; Medine, E.I.; Yavuz, M.; Unak, P.; Timur, S. Biological affinity evaluation of Lawsonia inermis origin Lawsone compound and its radioiodinated form via in vitro methods. J. Radioanal. Nucl. Chem. 2015, 303, 701–708. [Google Scholar] [CrossRef]
  35. Gan, C.; Zhao, Z.; Nan, D.D.; Yin, B.; Hu, J. Homoisoflavonoids as potential imaging agents for β-amyloid plaques in Alzheimer’s disease. Eur. J. Med. Chem. 2014, 76, 125–131. [Google Scholar] [CrossRef] [PubMed]
  36. Aras, O.; Takan, G.; Kilcar, A.Y.; Muftuler, F.Z.B. Extraction and radioiodination of Gingko flavonoids and monitoring the cellular incorporation. J. Radioanal. Nucl. Chem. 2016, 310, 271–278. [Google Scholar] [CrossRef]
  37. Jiang, C.; Gao, M.; Li, Y.; Huang, D.; Yao, N.; Ji, Y.; Liu, X.; Zhang, D.; Wang, X.; Yin, Z.; et al. Exploring diagnostic potentials of radioiodinated sennidin A in rat model of reperfused myocardial infarction. Int. J. Pharm. 2015, 495, 31–40. [Google Scholar] [CrossRef]
  38. Liu, X.; Feng, Y.; Jiang, C.; Lou, B.; Li, Y.; Liu, W.; Yao, N.; Gao, M.; Ji, Y.; Wang, Q.; et al. Radiopharmaceutical evaluation of (131)I-protohypericin as a necrosis avid compound. J. Drug Target. 2015, 23, 417–426. [Google Scholar] [CrossRef]
  39. Zhang, D.; Huang, D.; Ji, Y.; Jiang, C.; Li, Y.; Gao, M.; Yao, N.; Liu, X.; Shao, H.; Jing, S.; et al. Experimental evaluation of radioiodinated sennoside B as a necrosis-avid tracer agent. J. Drug Target. 2015, 23, 180–190. [Google Scholar] [CrossRef]
  40. Yang, H.L.; Chen, S.C.; Kumar, K.J.S.; Yu, K.N.; Chao, P.D.L.; Tsai, S.Y.; Hou, Y.C.; Hseu, Y.C. Antioxidant and anti-inflammatory potential of hesperetin metabolites obtained from hesperetin-administered rat serum: An ex vivo approach. J. Agric. Food Chem. 2012, 60, 522–532. [Google Scholar] [CrossRef]
  41. Shin, K.C.; Nam, H.K.; Oh, D.K. Hydrolysis of flavanone glycosides by β-glucosidase from Pyrococcus furiosus and its application to the production of flavanone aglycones from citrus extracts. J. Agric. Food Chem. 2013, 61, 11532–11540. [Google Scholar] [CrossRef]
  42. Jeon, J.; Ma, S.-Y.; Choi, D.; Kang, J.; Nam, Y.; Yoon, S.; Park, S. Radiosynthesis of 123I-labeled hesperetin for biodistribution study of orally administered hesperetin. J. Radioanal. Nucl. Chem. 2015, 306, 437–443. [Google Scholar] [CrossRef]
  43. Chua, L.S. A review on plant-based rutin extraction methods and its pharmacological activities. J. Ethnopharmacol. 2013, 150, 805–817. [Google Scholar] [CrossRef] [PubMed]
  44. Choi, M.H.; Rho, J.K.; Kang, J.A.; Shim, H.E.; Nam, Y.R.; Yoon, S.; Kim, H.R.; Choi, D.S.; Park, S.H.; Jang, B.-S.; et al. Efficient radiolabeling of rutin with 125I and biodistribution study of radiolabeled rutin. J. Radioanal. Nucl. Chem. 2016, 308, 477–483. [Google Scholar] [CrossRef]
  45. Antonisamy, P.; Agastian, P.; Kang, C.W.; Kim, N.S.; Kim, J.H. Anti-inflammatory activity of rhein isolated from the flowers of Cassia fistula L. and possible underlying mechanisms. Saudi J. Biol. Sci. 2019, 26, 96–104. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, D.; Jin, Q.; Ni, Y.; Zhang, J. Discovery of necrosis avidity of rhein and its applications in necrosis imaging. J. Drug Target. 2020, 28, 904–912. [Google Scholar] [CrossRef]
  47. Wang, Q.; Yang, S.; Jiang, C.; Li, J.; Wang, C.; Chen, L.; Jin, Q.; Song, S.; Feng, Y.; Ni, Y.; et al. Discovery of Radioiodinated Monomeric Anthraquinones as a Novel Class of Necrosis Avid Agents for Early Imaging of Necrotic Myocardium. Sci. Rep. 2016, 6, 21341. [Google Scholar] [CrossRef] [Green Version]
  48. Al-Sharif, I.; Remmal, A.; Aboussekhra, A. Eugenol triggers apoptosis in breast cancer cells through E2F1/survivin down-regulation. BMC Cancer 2013, 13, 600. [Google Scholar] [CrossRef] [Green Version]
  49. Vidhya, N.; Devaraj, S.N. Induction of apoptosis by eugenol in human breast cancer cells. Indian J. Exp Biol. 2011, 49, 871–878. [Google Scholar]
  50. Ghosh, R.; Ganapathy, M.; Alworth, W.L.; Chan, D.C.; Kumar, A.P. Combination of 2-methoxyestradiol (2-ME2) and eugenol for apoptosis induction synergistically in androgen independent prostate cancer cells. J. Steroid Biochem. Mol. Biol. 2009, 113, 25–35. [Google Scholar] [CrossRef]
  51. Dervis, E.; Kilcar, A.Y.; Medine, E.I.; Tekin, V.; Cetkin, B.; Uygur, E.; Muftuler, F.Z.B. In Vitro Incorporation of Radioiodinated Eugenol on Adenocarcinoma Cell Lines (Caco2, MCF7, and PC3). Cancer Biother. Radiopharm. 2017, 32, 75–81. [Google Scholar] [CrossRef]
  52. Gibellini, L.; Pinti, M.; Nasi, M.; Montagna, J.P.; De Biasi, S.; Roat, E.; Bertoncelli, L.; Cooper, E.L.; Cossarizza, A. Quercetin and cancer chemoprevention. Evid Based Complement. Altern. Med. 2011, 2011, 591356. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Lin, C.; Yu, Y.; Zhao, H.G.; Yang, A.; Yan, H.; Cui, Y. Combination of quercetin with radiotherapy enhances tumor radiosensitivity in vitro and in vivo. Radiother. Oncol. 2012, 104, 395–400. [Google Scholar] [CrossRef] [PubMed]
  54. Xie, Q.; Li, X.; Wang, G.; Hou, X.; Wang, Y.; Yu, H.; Qu, C.; Luo, S.; Cui, Y.; Xia, C.; et al. Preparation and evaluation of 131I-quercetin as a novel radiotherapy agent against dedifferentiated thyroid cancer. J. Radioanal. Nucl. Chem. 2017, 311, 1697–1708. [Google Scholar] [CrossRef]
  55. Palensina, G.; Rosita, L.; Sagala, Z. Isolation of Arbutin from Leaves and Fruits of Buni (Antidesma Bunius L. Spreng) As Tyrosinase Enzym Inhibitor. Bioinform. Biomed. Res. J. 2021, 4, 8–18. [Google Scholar] [CrossRef]
  56. Ebadollahi, S.H.; Pouramir, M.; Zabihi, E.; Golpour, M.; Aghajanpour-Mir, M. The Effect of Arbutin on The Expression of Tumor Suppressor P53, BAX/BCL-2 Ratio and Oxidative Stress Induced by Tert-Butyl Hydroperoxide in Fibroblast and LNcap Cell Lines. Cell J. 2021, 22, 532–541. [Google Scholar] [CrossRef]
  57. Huynh, P.T.; Ha, Y.S.; Lee, W.; Yoo, J. Radio-Iodinated arbutin for tumor imaging. J. Radiopharm. Mol. Probes 2017, 3, 72–79. [Google Scholar]
  58. Liang, J.; Sun, Z.; Zhang, D.; Jin, Q.; Cai, L.; Ma, L.; Liu, W.; Ni, Y.; Zhang, J.; Yin, Z. First Evaluation of Radioiodinated Flavonoids as Necrosis-Avid Agents and Application in Early Assessment of Tumor Necrosis. Mol. Pharm. 2018, 15, 207–215. [Google Scholar] [CrossRef]
  59. Chen, G.; Pi, X.M.; Yu, C.Y. A new naphthalenone isolated from the green walnut husks of Juglans mandshurica Maxim. Nat. Prod. Res. 2015, 29, 174–179. [Google Scholar] [CrossRef]
  60. Su, C.; Zhang, D.; Bao, N.; Ji, A.; Feng, Y.; Chen, L.; Ni, Y.; Zhang, J.; Yin, Z.-Q. Evaluation of Radioiodinated 1,4-Naphthoquinones as Necrosis Avid Agents for Rapid Myocardium Necrosis Imaging. Mol. Imaging Biol. 2017, 20, 74–84. [Google Scholar] [CrossRef]
  61. Chen, C.A.; Chang, H.H.; Kao, C.Y.; Tsai, T.H.; Chen, Y.J. Plumbagin, isolated from Plumbago zeylanica, induces cell death through apoptosis in human pancreatic cancer cells. Pancreatology 2009, 9, 797–809. [Google Scholar] [CrossRef]
  62. Aminin, D.; Polonik, S. 1,4-Naphthoquinones: Some Biological Properties and Application. Chem. Pharm. Bull. 2020, 68, 46–57. [Google Scholar] [CrossRef] [PubMed]
  63. Peñalver, P.; Belmonte-Reche, E.; Adán, N.; Caro, M.; Mateos-Martín, M.L.; Delgado, M.; González-Rey, E.; Morales, J.C. Alkylated resveratrol prodrugs and metabolites as potential therapeutics for neurodegenerative diseases. Eur. J. Med. Chem. 2018, 146, 123–138. [Google Scholar] [CrossRef]
  64. Jin, F.; Wu, Q.; Lu, Y.F.; Gong, Q.H.; Shi, J.S. Neuroprotective effect of resveratrol on 6-OHDA-induced Parkinson’s disease in rats. Eur. J. Pharmacol. 2008, 600, 78–82. [Google Scholar] [CrossRef]
  65. Guimón, J.; Guimón, P. How ready-to-use therapeutic food shapes a new technological regime to treat child malnutrition. Technol. Forecast. Soc. Chang. 2012, 79, 1319–1327. [Google Scholar] [CrossRef] [Green Version]
  66. Leis, K.; Baska, A.; Bereźnicka, W.; Marjańska, A.; Mazur, E.; Lewandowski, B.T.; Kałużny, K.; Gałązka, P. Resveratrol in the treatment of neuroblastoma: A review. Rev. Neurosci. 2020, 31, 873–881. [Google Scholar] [CrossRef] [PubMed]
  67. Karatay, K.B.; Kilcar, A.Y.; Guldu, O.K.; Medine, E.I.; Muftuler, F.Z.B. Isolation of resveratrol from peanut sprouts, radioiodination and investigation of its bioactivity on neuroblastoma cell lines. J. Radioanal. Nucl. Chem. 2020, 325, 75–84. [Google Scholar] [CrossRef]
  68. Sakai, T.; Kogiso, M. Soy isoflavones and immunity. J. Med. Investig. 2008, 55, 167–173. [Google Scholar] [CrossRef] [Green Version]
  69. Yuseran, H.; Hartoyo, E.; Nurseta, T.; Kalim, H. Molecular docking of genistein on estrogen receptors, promoter region of BCLX, caspase-3, Ki-67, cyclin D1, and telomere activity. J. Taibah Univ. Med. Sci. 2019, 14, 79–87. [Google Scholar] [CrossRef]
  70. Rajah, T.T.; Du, N.; Drews, N.; Cohn, R. Genistein in the presence of 17beta-estradiol inhibits proliferation of ERbeta breast cancer cells. Pharmacology 2009, 84, 68–73. [Google Scholar] [CrossRef]
  71. Ramdhani, D.; Widyasari, E.M.; Sriyani, M.E.; Arnanda, Q.P.; Watabe, H. Iodine-131 labeled genistein as a potential radiotracer for breast cancer. Heliyon 2020, 6, e04780. [Google Scholar] [CrossRef]
  72. Sp, N.; Kang, D.Y.; Lee, J.M.; Bae, S.W.; Jang, K.J. Potential Antitumor Effects of 6-Gingerol in p53-Dependent Mitochondrial Apoptosis and Inhibition of Tumor Sphere Formation in Breast Cancer Cells. Int. J. Mol. Sci. 2021, 22, 4660. [Google Scholar] [CrossRef] [PubMed]
  73. Ray, A.; Vasudevan, S.; Sengupta, S. 6-Shogaol Inhibits Breast Cancer Cells and Stem Cell-Like Spheroids by Modulation of Notch Signaling Pathway and Induction of Autophagic Cell Death. PLoS ONE 2015, 10, e0137614. [Google Scholar] [CrossRef] [PubMed]
  74. Karatay, K.B.; Kılçar, A.Y.; Derviş, E.; Müftüler, F.Z.B. Radioiodinated Ginger Compounds (6-gingerol and 6-shogaol) and Incorporation Assays on Breast Cancer Cells. Anticancer Agents Med. Chem. 2020, 20, 1129–1139. [Google Scholar] [CrossRef]
  75. İnce, İ.; Yıldırım, Y.; Güler, G.; Medine, E.İ.; Ballıca, G.; Kuşdemir, B.C.; Göker, E. Synthesis and characterization of folic acid-chitosan nanoparticles loaded with thymoquinone to target ovarian cancer cells. J. Radioanal. Nucl. Chem. 2020, 324, 71–85. [Google Scholar] [CrossRef]
  76. Woo, C.C.; Kumar, A.P.; Sethi, G.; Tan, K.H. Thymoquinone: Potential cure for inflammatory disorders and cancer. Biochem. Pharm. 2012, 83, 443–451. [Google Scholar] [CrossRef] [PubMed]
  77. Destito, G.; Yeh, R.; Rae, C.S.; Finn, M.G.; Manchester, M. Folic Acid-Mediated Targeting of Cowpea Mosaic Virus Particles to Tumor Cells. Chem. Biol. 2007, 14, 1152–1162. [Google Scholar] [CrossRef] [Green Version]
  78. Sriyani, M.E.; Nuraeni, W.; Rosyidiah, E.; Widyasari, E.M.; Saraswati, A.; Shintia, M. Quality control and stability study of the [131I]I-rutin produced in acidic condition. AIP Conf. Proc. 2021, 2381, 020082. [Google Scholar] [CrossRef]
  79. Kumar, N.; Pruthi, V. Potential applications of ferulic acid from natural sources. Biotechnol. Rep. 2014, 4, 86–93. [Google Scholar] [CrossRef] [Green Version]
  80. Eroğlu, C.; Seçme, M.; Bağcı, G.; Dodurga, Y. Assessment of the anticancer mechanism of ferulic acid via cell cycle and apoptotic pathways in human prostate cancer cell lines. Tumor Biol. 2015, 36, 9437–9446. [Google Scholar] [CrossRef]
  81. Zhang, X.D.; Wu, Q.; Yang, S.H. Ferulic acid promoting apoptosis in human osteosarcoma cell lines. Pak. J. Med. Sci. 2017, 33, 127–131. [Google Scholar] [CrossRef]
  82. Sedik, G.A.; Rizq, R.S.A.; Ibrahim, I.T.; Elzanfaly, E.S.; Motaleb, M.A. Miniaturized chromatographic systems for radiochemical purity evaluation of (131)I-Ferulic acid as a new candidate in nuclear medicine applications. Appl. Radiat. Isot. 2021, 167, 109370. [Google Scholar] [CrossRef] [PubMed]
  83. Selim, A.A.; Essa, B.M.; Abdelmonem, I.M.; Amin, M.A.; Sarhan, M.O. Extraction, purification and radioiodination of Khellin as cancer theranostic agent. Appl. Radiat. Isot. 2021, 178, 109970. [Google Scholar] [CrossRef] [PubMed]
  84. Hueza, I.M.; Raspantini, P.C.; Raspantini, L.E.; Latorre, A.O.; Górniak, S.L. Zearalenone, an estrogenic mycotoxin, is an immunotoxic compound. Toxins 2014, 6, 1080–1095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Zou, Z.; He, Z.; Li, H.; Han, P.; Meng, X.; Zhang, Y.; Zhou, F.; Ouyang, K.-P.; Chen, X.-Y.; Tang, J. In vitro removal of deoxynivalenol and T-2 toxin by lactic acid bacteria. Food Sci. Biotechnol. 2012, 21, 1677–1683. [Google Scholar] [CrossRef]
  86. Sanad, M.H.; Farag, A.B.; Bassem, S.A.; Marzook, F.A. Radioiodination of zearalenone and determination of Lactobacillus plantarum effect of on zearalenone organ distribution: In silico study and preclinical evaluation. Toxicol. Rep. 2022, 9, 470–479. [Google Scholar] [CrossRef]
  87. Subramanian, V.S.; Sabui, S.; Teafatiller, T.; Bohl, J.A.; Said, H.M. Structure/functional aspects of the human riboflavin transporter-3 (SLC52A3): Role of the predicted glycosylation and substrate-interacting sites. Am. J. Physiol.-Cell Physiol. 2017, 313, C228–C238. [Google Scholar] [CrossRef] [Green Version]
  88. Bulas, S.; Bedoukian, E.C.; O’Neil, E.C.; Krantz, I.D.; Yum, S.W.; Liu, G.T.; Aleman, T.S. Ocular Biomarkers of Riboflavin Transporter Deficiency. J. Neuroophthalmol. 2022. [Google Scholar] [CrossRef]
  89. Li, J.; Chen, Y.; Peng, C.; Hong, X.; Liu, H.; Fang, J.; Zhuang, R.; Pan, W.; Zhang, D.; Guo, Z.; et al. Micro-SPECT Imaging of Acute Ischemic Stroke with Radioiodinated Riboflavin in Rat MCAO Models via Riboflavin Transporter Targeting. ACS Chem. Neurosci. 2022, 13, 1966–1973. [Google Scholar] [CrossRef]
  90. Jeung, Y.J.; Kim, H.G.; Ahn, J.; Lee, H.J.; Lee, S.B.; Won, M.; Jung, C.R.; Im, J.Y.; Kim, B.K.; Park, S.K.; et al. Shikonin induces apoptosis of lung cancer cells via activation of FOXO3a/EGR1/SIRT1 signaling antagonized by p300. Biochim. Biophys. Acta 2016, 1863, 2584–2593. [Google Scholar] [CrossRef]
  91. Yeh, Y.-C.; Liu, T.-J.; Lai, H.-C. Shikonin Induces Apoptosis, Necrosis, and Premature Senescence of Human A549 Lung Cancer Cells through Upregulation of p53 Expression. Evid.-Based Complement. Altern. Med. 2015, 2015, 620383. [Google Scholar] [CrossRef] [Green Version]
  92. Selim, A.A.; Motaleb, M.A.; Fayez, H.A. Lung Cancer-Targeted [131I]-Iodoshikonin as Theranostic Agent: Radiolabeling, In Vivo Pharmacokinetics and Biodistribution. Pharm. Chem. J. 2022, 55, 1163–1168. [Google Scholar] [CrossRef]
  93. Mushtaq, S.; Jeon, J.; Shaheen, A.; Jang, B.S.; Park, S.H. Critical analysis of radioiodination techniques for micro and macro organic molecules. J. Radioanal. Nucl. Chem. 2016, 309, 859–889. [Google Scholar] [CrossRef]
  94. Takahashi, M.; Seki, K.-i.; Nishijima, K.-i.; Zhao, S.; Kuge, Y.; Tamaki, N.; Ohkura, K. Synthesis of a radioiodinated thymidine phosphorylase inhibitor and its preliminary evaluation as a potential SPECT tracer for angiogenic enzyme expression. J. Label. Compd. Radiopharm. 2008, 51, 384–387. [Google Scholar] [CrossRef]
  95. Zmuda, F.; Malviya, G.; Blair, A.; Boyd, M.; Chalmers, A.J.; Sutherland, A.; Pimlott, S.L. Synthesis and Evaluation of a Radioiodinated Tracer with Specificity for Poly(ADP-ribose) Polymerase-1 (PARP-1) in Vivo. J. Med. Chem. 2015, 58, 8683–8693. [Google Scholar] [CrossRef] [Green Version]
  96. Sadeghzadeh, M.; Daha, F.J.; Sheibani, S.; Erfani, M. Radioiodination of 4-benzyl-1-(3-iodobenzylsulfonyl)piperidine, 4-(3-iodobenzyl)-1-(benzylsulfonyl)piperazine and their derivatives via isotopic and non-isotopic exchange reactions. J. Radioanal. Nucl. Chem. 2014, 302, 1119–1125. [Google Scholar] [CrossRef]
  97. Hagimori, M.; Temma, T.; Kudo, S.; Sano, K.; Kondo, N.; Mukai, T. Synthesis of radioiodinated probes targeted toward matrix metalloproteinase-12. Bioorg. Med. Chem. Lett. 2018, 28, 193–195. [Google Scholar] [CrossRef]
  98. Lee, Y.-S. Radiopharmaceuticals for Molecular Imaging. Open Nucl. Med. J. 2010, 2, 178–185. [Google Scholar] [CrossRef]
  99. Waterhouse, R.N. Determination of lipophilicity and its use as a predictor of blood–brain barrier penetration of molecular imaging agents. Mol. Imaging Biol. 2003, 5, 376–389. [Google Scholar] [CrossRef]
  100. Arnott, J.; Lobo, S. The influence of lipophilicity in drug discovery and design. Expert Opin. Drug Discov. 2012, 7, 863–875. [Google Scholar] [CrossRef]
  101. Kratochwil, N.A.; Huber, W.; Müller, F.; Kansy, M.; Gerber, P.R. Predicting plasma protein binding of drugs: A new approach. Biochem. Pharm. 2002, 64, 1355–1374. [Google Scholar] [CrossRef]
  102. De Kruijff, R.M.; Wolterbeek, H.T.; Denkova, A.G. A Critical Review of Alpha Radionuclide Therapy—How to Deal with Recoiling Daughters? Pharmaceuticals 2015, 8, 321–336. [Google Scholar] [CrossRef] [PubMed]
  103. İlem-Özdemir, D.; Ekinci, M.; Gundogdu, E.; Asikoglu, M. Estimating Binding Capability of Radiopharmaceuticals by Cell Culture Studies. Int. J. Med. Nano Res. 2016, 3, 014. [Google Scholar] [CrossRef]
  104. Motaleb, M.A.; Ibrahim, I.T.; Sayyed, M.E.; Awad, G.A.S. (131)I-trazodone: Preparation, quality control and in vivo biodistribution study by intranasal and intravenous routes as a hopeful brain imaging radiopharmaceutical. Rev. Esp Med. Nucl. Imagen Mol. 2017, 36, 371–376. [Google Scholar] [CrossRef] [PubMed]
  105. Rebischung, C.; Hoffmann, D.; Stefani, L.; Desruet, M.D.; Wang, K.; Adelstein, S.J.; Artignan, X.; Vincent, F.; Gauchez, A.S.; Zhang, H.; et al. First human treatment of resistant neoplastic meningitis by intrathecal administration of MTX plus (125)IUdR. Int. J. Radiat. Biol. 2008, 84, 1123–1129. [Google Scholar] [CrossRef] [PubMed]
  106. Shuryak, I.; Dadachova, E. New Approaches for Modeling Radiopharmaceutical Pharmacokinetics Using Continuous Distributions of Rates. J. Nucl. Med. 2015, 56, 1622–1628. [Google Scholar] [CrossRef] [Green Version]
  107. Koziorowski, J.; Behe, M.; Decristoforo, C.; Ballinger, J.; Elsinga, P.; Ferrari, V.; Peitl, P.K.; Todde, S.; Mindt, T.L. Position paper on requirements for toxicological studies in the specific case of radiopharmaceuticals. EJNMMI Radiopharm. Chem. 2016, 1, 1. [Google Scholar] [CrossRef] [Green Version]
  108. Eberlein, U.; Bröer, J.H.; Vandevoorde, C.; Santos, P.; Bardiès, M.; Bacher, K.; Nosske, D.; Lassmann, M. Biokinetics and dosimetry of commonly used radiopharmaceuticals in diagnostic nuclear medicine—A review. Eur. J. Nucl. Med. Mol. Imaging 2011, 38, 2269–2281. [Google Scholar] [CrossRef] [Green Version]
  109. El-Tawoosy, M.; Ibrahim, I. Radioiodination and biological evaluation of salbutamol as a β2-adrenoceptor agonist. Radiochemistry 2012, 54, 401–406. [Google Scholar] [CrossRef]
  110. El-Azony, K.; El-Mohty, A.; Seddik, U.; Khater, s. Radioiodination and bioevaluation of nitrofurantoin for urinary tract imaging. J. Label. Compd. Radiopharm. 2012, 55, 315–319. [Google Scholar] [CrossRef]
  111. Kiess, A.P.; Minn, I.; Chen, Y.; Hobbs, R.; Sgouros, G.; Mease, R.C.; Pullambhatla, M.; Shen, C.J.; Foss, C.A.; Pomper, M.G. Auger Radiopharmaceutical Therapy Targeting Prostate-Specific Membrane Antigen. J. Nucl. Med. 2015, 56, 1401–1407. [Google Scholar] [CrossRef] [Green Version]
  112. Moustapha, M.E.; Motaleb, M.A.; Ibrahim, I.T.; Moustafa, M.E. Oxidative radioiodination of aripiprazole by chloramine-T as a route to a potential brain imaging agent: A mechanistic approach. Radiochemistry 2013, 55, 116–122. [Google Scholar] [CrossRef]
  113. Abdel-Bary, H.M.; Moustafa, K.A.; Abdel-Ghaney, I.Y.; Sallam, K.M.; Shamsel-Din, H.A. Synthesis and radioiodination of new dipeptide coupled with biologically active pyridine moiety. J. Radioanal. Nucl. Chem. 2013, 298, 9–18. [Google Scholar] [CrossRef]
  114. Amin, A.; Soliman, S.; El-Aziz, H. Preparation and biodistribution of [125I]Melphalan: A potential radioligand for diagnostic and therapeutic applications. J. Label. Compd. Radiopharm. 2009, 53, 1–5. [Google Scholar] [CrossRef]
  115. Amin, A.; Soliman, S.; El-Aziz, H.; El-Enein, S. Radioiodination of Zaleplon and Its in-vivo Biologic Behavior in Mice: An Imaging Probe for Brain. Int. J. Chem. 2013, 6, 17. [Google Scholar] [CrossRef]
  116. Avcıbaşı, U.; demiroğlu, H.; Unak, P.; Müftüler, F.; Ichedef, C.A.; Gumuser, F. In vivo biodistribution of 131 I labeled bleomycin (BLM) and isomers (A2 and B2) on experimental animal models. J. Radioanal. Nucl. Chem. 2010, 285, 207–214. [Google Scholar] [CrossRef]
  117. Baş, U.; Demiroğlu, H.; Ediz, M.; Akalın, H.; Şkan, E.; Senay, H.; Türkcan, C.; Ozcan, Y.; Akgöl, S.; Avcibaşi, N. Radiolabeling of new generation magnetic poly(HEMA-MAPA) nanoparticles with I-131 and preliminary investigation of its radiopharmaceutical potential using albino Wistar rats. J. Label. Compd. Radiopharm. 2013, 56, 708–716. [Google Scholar] [CrossRef]
  118. Hussien, H.; Goud, A.A.; Amin, A.M.; El-Sheikh, R.; Seddik, U. Comparative study between chloramine-T and iodogen to prepare radioiodinated etodolac for inflammation imaging. J. Radioanal. Nucl. Chem. 2011, 288, 9–15. [Google Scholar] [CrossRef]
  119. Wang, K.; Adelstein, S.J.; Kassis, A.I. DMSO increases radioiodination yield of radiopharmaceuticals. Appl. Radiat. Isot. 2008, 66, 50–59. [Google Scholar] [CrossRef] [Green Version]
  120. Li, G.; Kakarla, R.; Gerritz, S.W. A fast and efficient bromination of isoxazoles and pyrazoles by microwave irradiation. Tetrahedron Lett. 2007, 48, 4595–4599. [Google Scholar] [CrossRef]
  121. Mattner, F.; Mardon, K.; Katsifis, A. Pharmacological evaluation of [123I]-CLINDE: A radioiodinated imidazopyridine-3-acetamide for the study of peripheral benzodiazepine binding sites (PBBS). Eur. J. Nucl. Med. Mol. Imaging 2008, 35, 779–789. [Google Scholar] [CrossRef]
  122. Kim, A.; Choi, K.H. Purification System of 131I-Metaiodobenzylguanidine. In Proceedings of the Transactions of the Korean Nuclear Society Autumn Meeting 2019, Goyang, Republic of Korea, 24–25 October 2019. [Google Scholar]
  123. Chitneni, S.K.; Reitman, Z.J.; Spicehandler, R.; Gooden, D.M.; Yan, H.; Zalutsky, M.R. Synthesis and evaluation of radiolabeled AGI-5198 analogues as candidate radiotracers for imaging mutant IDH1 expression in tumors. Bioorg. Med. Chem. Lett. 2018, 28, 694–699. [Google Scholar] [CrossRef] [PubMed]
  124. Lemps, R.D.; Desruet, M.; Bacot, S.; Ahmadi, M.; Ghezzi, C.; Desruet, M.; Fagret, D.; Berger, F. Iodogen-mediated radiolabeling of Bevacizumab with I-123 for clinical applications. Eur. J. Nucl. Med. Mol. Imaging 2014, 41, S447. [Google Scholar]
  125. Visser, G.W.; Klok, R.P.; Gebbinck, J.W.; ter Linden, T.; van Dongen, G.A.; Molthoff, C.F. Optimal quality (131)I-monoclonal antibodies on high-dose labeling in a large reaction volume and temporarily coating the antibody with IODO-GEN. J. Nucl. Med. 2001, 42, 509–519. [Google Scholar] [PubMed]
  126. Molavipordanjani, S.; Tolmachev, V.; Hosseinimehr, S.J. Basic and practical concepts of radiopharmaceutical purification methods. Drug Discov. Today 2019, 24, 315–324. [Google Scholar] [CrossRef] [PubMed]
  127. Amin, A.; Farrag, N.; AbdEl-Bary, A. Iodine-125-Chlorambucil as Possible Radio Anticancer for Diagnosis and Therapy of Cancer: Preparation and Tissue Distribution. Br. J. Pharm. Res. 2014, 4, 1873–1885. [Google Scholar] [CrossRef]
  128. Spetz, J.; Rudqvist, N.; Forssell-Aronsson, E. Biodistribution and dosimetry of free 211At, 125I- and 131I- in rats. Cancer Biother. Radiopharm. 2013, 28, 657–664. [Google Scholar] [CrossRef] [Green Version]
  129. Cavina, L.; van der Born, D.; Klaren, P.H.M.; Feiters, M.C.; Boerman, O.C.; Rutjes, F. Design of Radioiodinated Pharmaceuticals: Structural Features Affecting Metabolic Stability towards in Vivo Deiodination. Eur. J. Org. Chem. 2017, 2017, 3387–3414. [Google Scholar] [CrossRef] [Green Version]
  130. Kim, E.J.; Kim, B.S.; Choi, D.B.; Chi, S.G.; Choi, T.H. Enhanced tumor retention of radioiodinated anti-epidermal growth factor receptor antibody using novel bifunctional iodination linker for radioimmunotherapy. Oncol. Rep. 2016, 35, 3159–3168. [Google Scholar] [CrossRef] [Green Version]
  131. Tong, R.; Kohane, D.S. New Strategies in Cancer Nanomedicine. Annu Rev. Pharm. Toxicol. 2016, 56, 41–57. [Google Scholar] [CrossRef]
Figure 1. Basic scheme of natural-based radiopharmaceuticals with iodine radioisotope.
Figure 1. Basic scheme of natural-based radiopharmaceuticals with iodine radioisotope.
Molecules 27 08009 g001
Figure 2. Development of radiopharmaceuticals from natural compounds in comparison to the conventional radiopharmaceuticals.
Figure 2. Development of radiopharmaceuticals from natural compounds in comparison to the conventional radiopharmaceuticals.
Molecules 27 08009 g002
Figure 3. Research of radiopharmaceuticals from natural compounds in last 10 years.
Figure 3. Research of radiopharmaceuticals from natural compounds in last 10 years.
Molecules 27 08009 g003
Figure 4. (a) Synthesis method and type of oxidizing agents used; (b) Purity of the 32 candidates of natural compound-based radiopharmaceuticals.
Figure 4. (a) Synthesis method and type of oxidizing agents used; (b) Purity of the 32 candidates of natural compound-based radiopharmaceuticals.
Molecules 27 08009 g004
Figure 5. Preclinical studies of radiopharmaceuticals.
Figure 5. Preclinical studies of radiopharmaceuticals.
Molecules 27 08009 g005
Figure 6. (a) Structure of ∆9-THC, 2-iodo-∆8-THC and 5′-iodo-∆8-THC; (b) The mechanism of deiodination reaction of 5′-iodo-∆8-THC; (c) Recommendation structure of iodinated THC which is stable against deiodination reaction [130].
Figure 6. (a) Structure of ∆9-THC, 2-iodo-∆8-THC and 5′-iodo-∆8-THC; (b) The mechanism of deiodination reaction of 5′-iodo-∆8-THC; (c) Recommendation structure of iodinated THC which is stable against deiodination reaction [130].
Molecules 27 08009 g006
Figure 7. A schematic diagram of [131I]FATQSCNPs.
Figure 7. A schematic diagram of [131I]FATQSCNPs.
Molecules 27 08009 g007
Figure 8. The literature search flow.
Figure 8. The literature search flow.
Molecules 27 08009 g008
Table 1. Radioisotopes of iodine.
Table 1. Radioisotopes of iodine.
RadioisotopeHalf LifeEmission TypeApplicationRefs.
iodine-12313.2 hGamma, EC 1/augerSPECT 3 diagnostic[8,9,10]
iodine-1244.8 daysPositronPET 2 diagnostic[11,12,13]
iodine-12560 daysGamma, EC 1/AugerPreclinical study, Radiotherapy SPECT 3 diagnostic[14,15]
iodine-1318.04 daysGamma, betaRadiotherapy, SPECT3 and PET 2 diagnostic[8,9,10]
1 EC: Electron capture, 2 PET: Positron emission tomography, 3 SPECT: Single photon emission tomography.
Table 2. Recent report of radiopharmaceuticals from natural compound with iodine radioisotopes and their importance of development/ application.
Table 2. Recent report of radiopharmaceuticals from natural compound with iodine radioisotopes and their importance of development/ application.
YearNatural CompoundsSourcesPharmacological
Activities
RadioisotopeApplicationRecent Research ReportedRefs.
2013Hydroxytyrosololive leaves extractAnticancer (breast, colon, prostate, and thyroid cancer)iodine-131Cancer therapyPreclinical study[21,22,23,24,25,26]
KhellinAmmi visnaga fruitsActivity against kidney disease and vitiligo, anticanceriodine-125Urinary tract imagingPreclinical study[27,28]
2014HypericinHypericum perforatum L.Antiviral, necrosis avidity and anticanceriodine-131Cancer therapyDosimetry prediction[29,30]
HypericinHypericum perforatum L.Antiviral, necrosis avidity and anticancer activityiodine-123Cancer therapyDosimetry prediction[30,31]
LawsoneLawsonia inermisAnticancer, antioxidant, and antibacterialiodine-131Cancer theranosticPreclinical study[32,33,34]
HomoisoflavonoidsHyacinthaceae and CaesalpinioideaeFormation, extension, and destabilization of Aβ aggregatesiodine-125diagnostic of b-amyloid plaques in Alzheimer’s diseasePreclinical study[35]
2015Gingko flavonoids (GFLAs)Egb761 extract of Gingko BilobaAnticanceriodine-131Cancer diagnosticCellular uptake[36]
Sinnidine ACassia Senna L.Structure similar to hypericin so it is predicted to have necrosis affinity like hypericiniodine-131Myocardial infarction imagingPreclinical study[37]
Protohypericin Hypericum perforatumStructure similar to hypericin so it is predicted to have necrosis affinity like hypericiniodine-131Cancer theranosticPreclinical study[38]
Sennoside BCassia senna L.Structure similar to hypericin so it is predicted to have necrosis affinity like hypericiniodine-131Necrosis-avid tracerPreclinical study[39]
Hesperetincitrus fruitsAnti-inflammatory, antioxidant, anticancer, antiviral, antiallergic, and neuroprotectiveiodine-123Radiotracer for some diseasePreclinical study[40,41,42]
2016Rutincitrus leavesAntitumor, cytotoxic, anti-inflammatory, antiestrogenic, antimicrobial, antiallergic, and antioxidantiodine-125Cancer diagnosticPreclinical study[43,44]
RheinCassia fistula L.Necrotic myocardiumiodine-131Myocardium necrosis imagingPreclinical study[45,46,47]
2017EugenolSyzygium aromaticumAnticancer (prostate, breast, colon, and cervical cancer)iodine-131Cancer therapyCellular uptake[48,49,50,51]
Quercetinvegetables, fruits, leaves, and grainsAnticanceriodine-131Thyroid cancer therapyPreclinical study[52,53,54]
Arbutinfresh fruit of the California buckeyeA tyrosinase inhibitor and antitumoriodine-131Tumor diagnosticPreclinical study[55,56,57]
2018VitexinPassiflora caerulea L.Necrosis-avid activityiodine-131Myocardium necrosis imagingPreclinical study[58]
Napthazarinegreen walnut husks of Juglans Mandshurica MaximNecrosis-avid activityiodine-131Myocardium necrosis imagingPreclinical study[59,60]
PlumbaginPlumbago zeylanicaNecrosis-avid activityiodine-131Myocardium necrosis imagingPreclinical study[60,61]
Jugloneleaves and nuts of various plants from the Juglandaceae familyNecrosis-avid activityiodine-131Myocardium necrosis imagingPreclinical study[60,62]
2019Resveratrolgrapes, peanut, and Polygonum cuspidatum rootAnti-inflammatory, antiapoptotic, neuroprotective antitumor, and immunological regulatoryiodine-131Neuroblastoma cells imagingCellular uptake[63,64,65,66,67]
2020GenisteinSoybeansAnticancer (Breast cancer)iodine-131Breast cancer diagnosticSynthesis[68,69,70,71]
6-Gingerolginger-roots extractAnticancer (breast cancer)iodine-131Breast cancer diagnosticCellular uptake[72,73,74]
6-Shogaolginger-roots extractAnticancer (breast cancer)iodine-131Breast cancer diagnosticCellular uptake[72,73,74]
ThymoquinoneNigella sativaAnticanceriodine-131Cancer theranosticCellular uptake[75]
FATQCSNPs (Folic acid-chitosan nanoparticles loaded with thymoquinone)Nigella sativaAnticanceriodine-131Cancer theranosticCellular uptake[75,76,77]
2021RutinSeveral fruits and vegetablesAnticanceriodine-131Cancer diagnosticPhysicochemical study[78]
Ferulic acidSeveral fruits and vegetablesAnticancer, antidiabetic, and activity against several neurodegenerative and cardiovascular diseasesiodine-131Cancer theranosticPreclinical study[79,80,81,82]
KhellinAmmi visnaga fruitsAnticanceriodine-131Cancer theranostic Preclinical study[83]
2022Zaeralenonecereal cropsAbility to bind competitively with estrogen receptorsiodine-125to study the the effect of Lactobacillus Plantarum on biodistribution pattern of ZaeralenonePreclinical study[84,85,86]
Riboflavinmeat, fish and fowl, eggs, dairy products, green vegetables, mushrooms, and almondsActivity against nervous system diseasesiodine-131Ischemic stroke diagnosticPreclinical study[87,88,89]
ShikoninLithospermum erythrorhizonAnticancer (lung cancer)iodine-131Lung cancer diagnosticPreclinical study[90,91,92]
Table 3. Characterization of natural compounds and synthesis of radiopharmaceuticals candidates from natural compounds using iodine radioisotope that were reported in the last 10 years (2013–2022).
Table 3. Characterization of natural compounds and synthesis of radiopharmaceuticals candidates from natural compounds using iodine radioisotope that were reported in the last 10 years (2013–2022).
Natural CompoundCharacterizationSynthesisIodinated Natural CompoundCharacterizationRefs.
Hydroxytyrosol
Molecules 27 08009 i001
LC-MS (liquid
chromatography-mass spectrometry) with positive mode [M+H] showed m/z 155.
iodogen[131I]hydroxytyrosol
Molecules 27 08009 i002
Structure was characterized by 1H NMR and 13C NMR Radiochemical Purity > 95% (by TLRC)[26]
Khellin
Molecules 27 08009 i003
Not reportedchloramine-T[125I]khellin
Molecules 27 08009 i004
Radiochemical Purity < 95% (by TLRC)[28]
Hypericin
Molecules 27 08009 i005
HPLC-UV with retention time of 7.85 miniodogen[131I]hypericin
Molecules 27 08009 i006
HPLC with retention time of 11.57 min
Radiochemical Purity: >95% (by HPLC)
[29,30]
Hypericin
Molecules 27 08009 i007
HPLC-UV with retention time of 7.85 miniodogen[123I]hypericin
Molecules 27 08009 i008
HPLC with retention time of 11.57 min
Radiochemical Purity: >95% (by HPLC)
[30,31]
Lawsone
Molecules 27 08009 i009
Structure was characterized by 1H NMR and 13C NMR iodogen[131I]lawsone
Molecules 27 08009 i010
Structure was characterized by 1H NMR and 13C NMR
Radiochemical Purity: <95% (by TLRC)
[33]
Homoisoflavonoid
Molecules 27 08009 i011
Structure was characterized by 1H NMR and 13C NMR iododestannylation[125I]I-Homoisoflavonoid
Molecules 27 08009 i012
Structure was characterized by 1H NMR and 13C NMR
-Radiochemical Purity: >95% (by HPLC)
[34]
GFLASCharacterized by HPLCiodogen[131I]GFLAS
Predicted structure have not reported
Radiochemical Purity: <95% (by TLRC)[36]
Sennidin A
Molecules 27 08009 i013
HPLC-UV with retention time of 9.98 miniodogen[131I]sennidin A
Molecules 27 08009 i014
HPLC-UV with a retention time of 11.76 min Radiochemical Purity: <95%[37]
Protohypericin
Molecules 27 08009 i015
HPLC-MS/MS [M,H]- with m/z 505
Structure was characterized by 1H NMR and 13C NMR
iodogen[131I]protohypericin
Molecules 27 08009 i016
Radiochemical Purity: >95% (by HPLC)[38]
Sennoside B
Molecules 27 08009 i017
HPLC with retention time of 7.09 miniodogen[131I]sennoside B
Molecules 27 08009 i018
HPLC with retention time 9.55 min Radiochemical Purity: >95% (by HPLC)[3]
Hesperetin
Molecules 27 08009 i019
LC/MS with [M,H]+ show m/z of 427
Structure was characterized by NMR
peracetic acid[123I]hesperetin
Molecules 27 08009 i020
Structure characterized by NMR and COSY analysis Radiochemical Purity: >95% (by HPLC)[42]
Rutin
Molecules 27 08009 i021
Structure was characterized by NMRchloramine-T[125I]rutin
Molecules 27 08009 i022
Structure was characterized by NMR
LC MS [M+H]+ with m/z of 737
Radiochemical Purity: >95% (by HPLC)
[44]
Rhein
Molecules 27 08009 i023
Not reportedperacetic acid[131I]rhein
Molecules 27 08009 i024
Structure was characterized by NMR
LC MS [M-H]- with m/z of 408.9
Radiochemical Purity: >95% (by HPLC)
[47]
Eugenol
Molecules 27 08009 i025
LC MS [M+H]+ with m/z of 164.80
HPLC with retention time of 12.456 min
iodogen[131I]eugenol
Molecules 27 08009 i026
Structure was characterized by NMR
Radiochemical Purity: >95% (by TLRC)
[51]
Quercetin
Molecules 27 08009 i027
Not reportedchloramine-T[131I]quercetin
Molecules 27 08009 i028
LC/MS characterization
Radiochemical Purity: >95% (by HPLC)
[54]
Arbutin
Molecules 27 08009 i029
HPLC with retention time of 1.6 minchloramine-T[131I]arbutin
Molecules 27 08009 i030
HPLC with retention time of 19,9 min Radiochemical Purity: >95%[57]
Vitexin
Molecules 27 08009 i031
Not reportediodogen[131I]vitexin
Molecules 27 08009 i032
Stucture was characterized by NMR
Radiochemical Purity: >95% (by HPLC)
[58]
Napthazarine
Molecules 27 08009 i033
Not reportediodogen[131I]napthazarine
Molecules 27 08009 i034
HPLC with retention time of 8.53 min Radiochemical Purity: >95% (by HPLC)[60]
Plumbagin
Molecules 27 08009 i035
Not reportediodogen[131I]plumbagin
Molecules 27 08009 i036
Radiochemical Purity: >95% (by TLRC)[60]
Juglone
Molecules 27 08009 i037
Not reportediodogen[131I]juglone
Molecules 27 08009 i038
Radiochemical Purity: >95% (by TLRC)[60]
Resveratrol
Molecules 27 08009 i039
Structure was characterized by NMR
LC MS [M+H]+ with m/z of 229.09
iodogen[131I]resveratrol
Molecules 27 08009 i040
Structure was characterized by NMR Radiochemical Purity: >95% (by TLRC)[67]
GenisteinNot reportedchloramine-T[131I]genistein
Predicted structure have not reported
Not reported Radiochemical Purity: >95% (by TLRC)[71]
6-GingerolNot reportediodogen[131I]6-gingerol
Predicted structure have not reported
Radiochemical Purity: >95% by TLRC)[74]
6-ShogaolNot reportediodogen[131I]6-shogaol
Predicted structure have not reported
Radiochemical Purity: >95% (by TLRC)[74]
ThymoquinoneCharacterized by FTIR has C-H (2950–2800 cm−1), C=C aromatic (1625–1440 cm−1) and C=O ketones (1700–1665 cm−1)iodogen[131I]thymoquinone
Predicted structure have not reported
Radiochemical Purity: <95% (by TLRC)[75]
FATQCSNPs Characterized by FTIR has amine stretch in Chitosan (3550–3250cm−1), OH from Folic acid (3200–2500 cm−1), C = O ketones from thymoquinone (1690 cm−1), C = O carboylic acid from Folic acid (1715 cm−1), C-C (1300–1100 cm−1) and C-O (1320–1210 cm−1)iodogen[131I]FATQCSNPs
Predicted structure have not reported
Radiochemical Purity < 95% (by TLRC)[75]
RutinNot reportedchloramine-T[131I]rutinRadiochemical Purity: <95% (by TLRC)[78]
Ferulic acid
Molecules 27 08009 i041
Not reportedchloramine-T[131I]ferulic acid
Molecules 27 08009 i042
LC/MS showed m.z 321.02
HPLC with retention time 17 min
Radiochemical Purity: >95%
[82]
Khellin
Molecules 27 08009 i043
Not reportediodogen[131I]khellin
Molecules 27 08009 i044
Radiochemical Purity: >95% (by HPLC)[83]
Zearalenone
Molecules 27 08009 i045
HPLC with retention time of 14.7 minchloramine-T[125I]I-zearalenone
Molecules 27 08009 i046
HPLC with retention time of 15.8 min
Purity: >95% (by HPLC)
[86]
Riboflavin
Molecules 27 08009 i047
Not reportediodogen[131I]I-riboflavin
Molecules 27 08009 i048
Radiochemical Purity: >95% (by paper chromatography)[89]
Shikonin
Molecules 27 08009 i049
Not reportedchloramine-T[131I]shikonin
Molecules 27 08009 i050
Structure was characterized by NMR
HPLC Rt of 8.14
Radiochemical Purity: <95%
[92]
Table 4. Evaluation and characterization of natural compound-based radiopharmaceutical candidates against with iodine radioisotopes.
Table 4. Evaluation and characterization of natural compound-based radiopharmaceutical candidates against with iodine radioisotopes.
CompoundStabilityLog PCell UptakePreclinical StudyDosimetryRefs.
[131I]hydroxytyrosol<4 h−0.41 ± 0.12Cellular uptake on Hutu80 (37.10%) > Caco2 (27.80%) > MCF7 (14.9%) > PC3 (14.50%)Biodistribution: highest uptake in bladder, stomach, and intestine.Not reported[26]
[125I]khellin>24 hNot reportedNot reportedBiodistribution: The highest uptake in heart, lung, and spleen.Not reported[28]
[131I]hypericinNot reportedNot reportedNot reportedBiodistribution: low uptake in necrosis cells but higher in lung, spleen, liverHigh absorbed radiation dose in necrotic tissues.[29,30]
[123I]hypericinNot reportedNot reportedNot reportedBiodistribution: high uptake in necrosis cells but lower in lung, spleen, liver.High absorbed radiation dose in necrotic tissues.[30,31]
[131I]lawsone<4 h−0.26 ± 0.06Keratinoccyte (25.46%) > BJ (5.43%) > MCF7 (5.32%) > Caco2) (5.28%) on 4 hBiodistribution: highest uptake in uterus, breast and ovary (female mice); and prostate (male mice)Not reported[33]
[125I]homoisoflavonoidsNot reportedNot reportedNot reportedBiodistribution in normal mice: high uptake in the brain with rapid clearance from the brain.Not reported[34]
[131I]GFLAS>24 h−0.99 ± 0.03Cellular uptake on PC3 > MCF7 not reportedNot reported[36]
[131I]sennidin AIn vivo stability > 48 h−1.11 ± 0.02Not reportedPharmacokinetics: AUC of 634.65 MBq/Lxh, clearance 0.02 L/h/kg. The elimination half-life (t1/2) of 11.75 hours
SPECT/CT image shows high accumulation of radioactivity in necrotic tissue.
Biodistribution: high uptake in necrotic tissues, liver, spleen and kidney
Not reported[37]
[131I]protohypericinNot reportedNot reportedNot reportedBiodistribution: the highest ratio of target/non-target tissues was 11.7 Pharmacokinetics: concentration after injection in blood 99.451±4.442 MBq/L t1/22 was 14.9 h using noncompartmental analyses (show fast blood clearance)
SPECT-CT, autoradiography, and histological staining showed high uptake in necrotic tissues
Not reported[38]
[131I]sennoside BNot reportedNot reportedNot reportedSPECT-CT showed selective accumulation of radioactivity in the necrotic tissues.
The highest biodistribution: the highest uptake in necrotic liver, necrotic muscle and kidney
Pharmacokinetics t1/2 8.6 h (fast clearance from blood)
Not reported[3]
[123I]hesperetin<4 hNot reportedNot reportedThe highest Biodistribution: highest uptake in stomach and intestine.Not reported[42]
[125I]rutinNot reportedNot reportedNot reportedBiodistribution and SPECT/CT studies in mice
oral administration: high biodistribution uptake in stomach and small intestine
intravena administration: highest biodistribution uptake in liver and small intestine
Not reported[44]
[131I]rhein>24 hNot reportedNot reportedStability > 24 h
Pharmacokinetics: t1/2 8.2 ± 0.49 h
Biodistribution: has optimum heart-to-blood, heart-to-liver and heart-to-lung ratios.
Not reported[47]
[131I]eugenolIn vivo stability > 48 h−1.50 ± 0.15In 4 h, cellular uptake on PC3 (54.35%)> MCF7 (45.68%)> Caco-2 (36.60%)Not reportedNot reported[51]
[131I]quercetin Not reportedCellular uptake in human thyroid: TT cell lines> FTC-133 cell lines> DRO cell lines Cells viability study with CCK-8 assay showed the rate of proliferation inhibiton of [131I]I-qQuercetin ≥ [131I+]qQuercetin > qQuercetin > iodine-131131IBiodistribution: the highest biodistribution uptake in tumors.
In vivo therapeutic efficacy study in tumors showed that a single dose can suppressed suppress tumor growth with mild side effects.
Not reported[54]
[131I]arbutin Not reportedNot reportedThe biodistribution study in CT26 tumor model mice were showed the highest uptake in bladder and kidneyNot reported[57]
[131I]vitexin 1.48 ± 0.06Not reportedPharmacokinetics: t1/2 5.3 h
Biodistribution: necrotic-viable myocardium ratio of 5.0 ± 0.9
SPECT/CT: clear necrosis imaging on CA4P-treated W256 tumors.
In vivo blocking study: could be blocked 51.95% and 64.29% by EB and cold vitexin
Not reported[58]
[131I]napthazarin Not reportedNot reportedBiodistribution: high necrotic-to-viable ratio and necrosis-to-blood ratio
Pharmacokinetic: t1/2 4.73 h
SPECT/CT: necrotic myocardium could be clearly visualized
in vitro DNA-binding: napthazarin could bind to DNA through intercalation
in vivo blocking study: necrotic muscle could be significantly blocked by excessive ethidium bromide (a typical DNA intercalator) and cold naphthazarin with 63.49 and 71.96% decline.
Not reported[60]
[131I]plumbagin>12 hNot reportedNot reportedBiodistribution: exhibited higher DNA-binding 5.60 × 104 M−1Not reported[60]
[131I]juglone>12 hNot reportedNot reportedBiodistribution: exhibited higher DNA-binding: 7.53 × 104 M−1Not reported[60]
[131I]resveratrol>24 h0.48 ± 0.2Cellular uptake on human neuroblastoma cell lines SK-N-AS (24.24%)> SH-SY5Y (15.04%) Not reportedNot reported[67]
[131I]genistein Evaluation have not reported [71]
[131I]6-gingerolNot reportedNot reportedCellular uptake in breast cancer cell lines MDA-MB-231: [131I]-6-sShogaol > [131I]-6-gGingerolNot reportedNot reported[74]
[131I]6-shogaolNot reportedNot reportedCellular uptake in breast cancer cell lines MCF7: [131I]-6-sShogaol similar to [131I]-6-g-GingerolNot reportedNot reported[74]
[131I]thymoquinone4 hNot reportedCellular uptake: SKOV3 (7.3%) > Caco-2 (5.75%) (in dose 200–1000 ng/mL)Not reportedNot reported[75]
[131I]FATQCSNPs 4 hNot reportedCellular uptake: SKOV3 (12.38%) > Caco-2 (6.73%) (in dose 200–1000 ng/mL)Not reportedNot reported[75]
[131I]rutinNot reported0.44 ± 0.16Not reportedNot reportedNot reported[78]
[131I]ferulic acid>24 hNot reportedNot reportedBiodistribution: %ID/gram in tumor s 4.35 ± 0.41 with tumor to muscle ratio 2.79Not reported[82]
[131I]khellin>24 hNot reportedNot reportedBiodistribution: the highest uptake in kidney, liver, intestine, tumorNot reported[83]
[125I]zearalenone>24 hNot reportedNot reportedBiodistribution in normal and bearing acid lactic mice showed a high accumulation in blood, liver, kidney, and intestineNot reported[86]
[131I]riboflavinNot reportedNot reportedNot reportedSPECT/CT image: uptake in the cerebral injury> normal brain
Autoradiography: infarcted to normal brain ratio 3.63
Blocking study: infarcted to normal brain ratio decrease to 1.98 after blocking
Not reported[89]
[131I]shikoninNot reportedNot reportedNot reportedBiodistribution the highest uptake in lung tissue (81.28% ID/g)
Pharmacokinetics: t1/2 elimination 40.05 ± 3.02 min.
Not reported[92]
Table 5. Challenges in the development of natural product-based radiopharmaceuticals.
Table 5. Challenges in the development of natural product-based radiopharmaceuticals.
NoChallengesCases on Previous Studies
1.Problem related to radiochemical
purity
Radiolabeled compounds have low radiochemical purity (RCP < 95%): [125I]khellin, [131I]lawsone, [131I]GFLAS, [131I]sennidin A, [131I]thymoquinone, [131I]FATQCSNPs, [131I]rutin, and [131I]shikonin
2.Problem related to biodistributionThe biodistribution pattern was high in certain organs, especially the thyroid, intestine and stomach: [131I]hydroxytyrosol, [123I]hesperetin, [125I]rutin, [131I]khellin, and [125I]zearalenone
Table 6. Critical points and considerations of radioiodination reaction method as a strategy for increasing the radiochemical purity.
Table 6. Critical points and considerations of radioiodination reaction method as a strategy for increasing the radiochemical purity.
Radioiodination MethodCritical Point that Needs to Be OptimizedConsiderationsRefs.
Electrophilic substitution
Chloramine-T (CAT)
  • pH
  • CAT concentration
  • Temperature
pH should be neutral, weak acid, or weak basic media.[28,109,110,111,112,113,114,115]
Excessive concentration causes oxidative side reactions such as polymerization, chlorination, and denaturation of the substrate.
Temperature to achieve the energy required for substitute H+ from the aromatic ring with radioactive iodonium ion.
Iodogen
  • pH
  • Iodogen concentration
  • Solvent
pH should be 7–8[116,117,118,119]
excessive concentration causes precipitates on the walls of the reaction vessel causing a low radiochemical purity.
Solvent: substrate in DMSO solvent showed with higher radiochemical purity RCP than substrate in aqueous solvent.
N-halosuccinimides (N-chlorosuccinimide and N-iodosuccinimide)
  • pH
  • Mediators
pH: N-iodosuccinimide with high activity in a strong acid medium[93,120,121]
Mediators such as NGA or mAB
Nucleophilic Substitutions (halogen and isotopic exchange)
  • Temperature
  • Reaction time
High temperature is required[93]
Reaction time: reactions take a long reaction time
Table 7. The resistance of some radioiodinated groups against in vivo deiodination [129].
Table 7. The resistance of some radioiodinated groups against in vivo deiodination [129].
Resistant to DeiodinationNon-Resistant to Deiodination
Iodinated carbon sp2Iodinated carbon sp and sp3
Iodoarenes Iodoaniline
Iodovinyl Iodophenols
IodoallylRadioiodinated nitrogen-containing (quinozalines, indoles, or imidazoles), and sulfur-containing (thiophenes) heterocycles
Radioiodinated oxygen-containing heterocycles
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Nurhidayah, W.; Setyawati, L.U.; Daruwati, I.; Gazzali, A.M.; Subroto, T.; Muchtaridi, M. Future Prospective of Radiopharmaceuticals from Natural Compounds Using Iodine Radioisotopes as Theranostic Agents. Molecules 2022, 27, 8009. https://doi.org/10.3390/molecules27228009

AMA Style

Nurhidayah W, Setyawati LU, Daruwati I, Gazzali AM, Subroto T, Muchtaridi M. Future Prospective of Radiopharmaceuticals from Natural Compounds Using Iodine Radioisotopes as Theranostic Agents. Molecules. 2022; 27(22):8009. https://doi.org/10.3390/molecules27228009

Chicago/Turabian Style

Nurhidayah, Wiwit, Luthfi Utami Setyawati, Isti Daruwati, Amirah Mohd Gazzali, Toto Subroto, and Muchtaridi Muchtaridi. 2022. "Future Prospective of Radiopharmaceuticals from Natural Compounds Using Iodine Radioisotopes as Theranostic Agents" Molecules 27, no. 22: 8009. https://doi.org/10.3390/molecules27228009

Article Metrics

Back to TopTop